Vous êtes sur la page 1sur 18

Physical Modeling of Seismic Soil-Pile-Structure

Interaction for Buildings on Soft Soils


Aslan S. Hokmabadi1; Behzad Fatahi, Ph.D.2; and Bijan Samali, D.Sc.3
Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Abstract: The present research intends to study the effects of the seismic soil-pile-structure interaction (SSPSI) on the dynamic response of
buildings with various heights by conducting a series of shaking table tests on 5-, 10-story, and 15-story model structures. Two types of foun-
dations for each case are investigated, including (1) a fixed-base structure, representing the situation excluding the soil-structure interaction; and
(2) a structure supported by an end-bearing pile foundation in soft soil. An advanced laminar soil container has been designed that uses three-
dimensional numerical modeling to minimize the boundary effects and to simulate free-field motion during the shaking table tests. Four real
earthquake events, including Kobe 1995, Northridge 1994, El Centro 1940, and Hachinohe 1968, are imposed to each model. According to the
experimental measurements, it is observed that the SSPSI amplifies the maximum lateral deflections and in turn interstory drifts of the structures
supported by end-bearing pile foundations in comparison with the fixed-base structures. The rocking component plays an important role in
increasing the lateral deflection of the superstructures, which can shift the performance level of the structures to near collapse or even collapse
levels and as a result should be assessed precisely in the seismic design of buildings resting on soft soils. DOI: 10.1061/(ASCE)GM.1943-
5622.0000396. © 2014 American Society of Civil Engineers.
Author keywords: Soil-pile-structure interaction; Seismic response; Shaking table tests; Laminar soil container; End-bearing pile foundations;
Moment-resisting buildings.

Introduction is particularly important in tall, slender, and closely spaced structures


that can be subjected to pounding when relative displacements are
The role of the seismic soil-pile-structure interaction (SSPSI) is usually large (Kramer 1996). Moreover, increase in the total deformation of
considered beneficial to the structural system under seismic loading the structure, and in turn secondary P-D effect, influences the total
because it elongates the period of the structure and increases the stability of the structure as well as the performance of the non-
damping of the structural system. Thus, consideration of the SSPSI structural elements. The lessons learned from postseismic obser-
tends to reduce the base shear, and in turn the structural demand, of vations of past earthquakes such as the 1985 Mexico City, 1994
the superstructure in comparison with the fixed-base condition. In Northridge, and 1995 Kobe earthquakes provided sufficient reasons
contrast, as shown in Fig. 1, the SSPSI may increase the overall to believe that the SSPSI effects should be investigated with greater
displacement of the superstructure in comparison with the fixed-base rigor and precision (Mendoza and Romo 1989; Mizuno et al. 1996;
condition because of translation and rotation of the foundation (e.g., Phanikanth et al. 2013).
Guin and Banerjee 1998; Han 2002). The rocking stiffness develops Performance-based seismic design is a modern approach to
as a result of the resistance of the piles to the vertical movement (Finn earthquake-resistant design that changes the emphasis from strength
2005), as shown in Fig. 1(b). Ma et al. (2009) mentioned that rocking to performance. Seismic performance (specifically, the performance
may be the most critical mode of vibration for a foundation, which level) is described by considering the maximum allowable damage
will lead to high motion amplitude when the excitation frequencies state (damage performance) for an identified seismic hazard (specif-
are near the resonance state. The increase in the lateral deformation ically, the hazard level). Performance levels describe the state of
of the building can change the performance level of the structure and structures after being subjected to a certain hazard level, and based on
FEMA 273=274 [Building Seismic Safety Council (BSSC) 1997]
they are classified as fully operational, operational, life safe, near
collapse, or collapse. Overall lateral deflection, ductility demand, and
1
Ph.D. Candidate, Centre for Built Infrastructure Research, School of interstory drifts are the most commonly used damage parameters.
Civil and Environmental Engineering, Univ. of Technology Sydney (UTS), Moreover, most of the force-based design codes use an additional
Sydney, NSW 2007, Australia (corresponding author). E-mail: aslan check in terms of limiting interstory drifts to ensure that particular
.sadeghihokmabadi@student.uts.edu.au deformation-based criteria are met. For example, ASCE 7-10 (ASCE
2
Senior Lecturer, Centre for Built Infrastructure Research, School of 2010) defines allowable story drift for structures by considering the
Civil and Environmental Engineering, Univ. of Technology Sydney (UTS), type and risk category of the structure. As a result, the SSPSI can alter
Sydney, NSW 2007, Australia. E-mail: behzad.fatahi@uts.edu.au a superstructure’s performance by influencing the dynamic properties
3
Professor, Centre for Built Infrastructure Research, School of Civil and of the structure. This phenomenon is investigated experimentally in
Environmental Engineering, Univ. of Technology Sydney (UTS), Sydney,
this article.
NSW 2007, Australia. E-mail: bijan.samali@uts.edu.au
Note. This manuscript was submitted on June 24, 2013; approved on Model tests in geotechnical engineering offer the advantage of
February 25, 2014; published online on April 11, 2014. Discussion period simulating complex systems under controlled conditions, pro-
open until September 11, 2014; separate discussions must be submitted for viding the opportunity to better understand the fundamental
individual papers. This paper is part of the International Journal of mechanisms of these systems. Such tests are often used as cali-
Geomechanics, © ASCE, ISSN 1532-3641/04014046(18)/$25.00. bration benchmarks for numerical or analytical methods, or to

© ASCE 04014046-1 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Schematic modeling of the 5-story multiple degree of freedom structure considering: (a) structure supported by end-bearing pile foundation
using foundation springs; (b) lateral deformation and rocking of the structure supported by end-bearing pile foundation; (c) lateral deformation of the
fixed-base structure

make quantitative predictions of the prototype response (Rayhani Although a number of single gravity (1g) model tests dealing
and El Naggar 2008). Shaking table tests are considered as 1g with the SSPSI effects on the seismic response of structures are
modeling, in which the gravity acceleration of the model and available in the literature as summarized in Table 1, most simplify
prototype stays the same. The shaking table test is relatively the superstructure as a single degree of freedom (SDOF) oscil-
cheap and easy to use in modeling complex prototypes, although lator in which the behavior of the soil-structure system may not
there is a lack of accuracy because of the 1g manner (e.g., dif- completely conform to reality and the effect of higher modes
ficulties in satisfying the scaling rules and low confining pressure would not be captured. In the current model tests, unlike the
of model affect test results especially in sandy soils). On the other previous efforts, a multistory frame for the superstructure is
hand, by increasing the gravity force via rotating the model in adopted, representing the dynamic properties of the prototype
centrifuge tests, it is possible to accurately model the soil stress- structure such as the natural frequency of the first and higher
strain condition as it exists in the prototype. In comparison, al- modes, number of stories, and the mass of the structure. A
though the centrifuge test models the stress-strain conditions complete set of experimental tests with various heights (5-, 10-,
accurately, it is difficult to build complex prototypes, and because and 15-story) and foundation types (fixed base and pile) was
of the small size of the model fewer instruments can be installed conducted, leading to experimentally comparable results used to
(Jakrapiyanun 2002). determine the influence of the SSPSI on the seismic response of
Scale models meet the requirements of similitude to the prototype superstructures. Moreover, an advanced laminar soil container
to differing degrees, and researchers may apply nomenclature such as has been designed using three-dimensional (3D) numerical
true, adequate, or distorted to the model (Moncarz and Krawinkler modeling to minimize the boundary effects and simulate the free-
1981). A true model fulfils all the similitude requirements. An ade- field motion during the shaking table tests.
quate model correctly scales the primary features of the problem, with The experimental model tests in this study have been carried out
secondary influences allowed to deviate while the prediction equation using the shaking table facilities located at the structures laboratory
is not significantly affected. Distorted models refer to those cases in of the University of Technology Sydney (UTS). The size of the
which deviation from similitude requirements distorts the prediction shaking table is 3 3 3 m, with a maximum payload of 9 Mg and
equation, or where compensating distortions in other dimensionless overturning moment of 100 kN×m. Furthermore, the shaking table
products are introduced to preserve the prediction equation. Meymand can apply a testing frequency range of 0.1–100 Hz and maximum
(1998) and Moss et al. (2010) explained that no governing equation acceleration of 624:5 m=s2 (62:5g) with no payload and
can be written describing the entire soil-structure system, nor can 68:82 m=s2 (60:9g) with 9 Mg of payload.
dimensional analysis or similitude theory be directly applied to this
complex system to achieve true model similarity. Thus, the viable Prototype Models
scale modeling approach for application of scale model similitude
consists of identifying and successfully modeling the primary forces Three moment-resisting building frames with total heights of 15 m
and processes in the system, while suppressing secondary effects, (5-story), 30 m (10-story), and 45 m (15-story), representing con-
thereby yielding an adequate model. ventional types of moment-resisting buildings, were selected for

© ASCE 04014046-2 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Table 1. Shaking Table Tests Previously Performed on Soil-Structure length of 1.2 and 30 m, respectively, giving the L=D ratio of 25. The
Systems Using Various Types of Soil Containers center-to-center spacing of the piles was 4 m (3:3D), which is in
Container type Reference Comment a reasonable range as adopted by other researchers (e.g., Shelke and
Patra 2008; Small and Zhang 2002). Because the superstructures consist
Rigid Gohl and Finn (1987) Adopted structural model: of three spans with a total width of 12 m, by adopting the aforemen-
Yan and Byrne (1989) hollow aluminum tubing, tioned pile setup and allocating one pile under each column, an efficient
Valsangkar et al. (1991) hollow aluminum piles, and pile foundation system to carry the applied structural loads was
Zen et al. (1992) retaining wall. designed. The piles were embedded into the bedrock, representing
Sato et al. (1995) Adopted soil model: dry typical end-bearing pile foundations as shown in Fig. 2.
Bathurst et al. (2007) Ottawa sand, saturated sand, To achieve a reasonable scale model, a dynamic similitude between
Ha et al. (2011) and saturated sand mixed the model and the prototype should be applied as described in the lit-
Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

with treated soil. erature (e.g., Langhaar 1951; Harris and Sabnis 1999; Meymand 1998).
Flexible Stanton et al. (1998) Adopted structural model: Dynamic similitude governs a condition where homologous parts of
Richards et al. (1990) steel piles only, wooden the model and prototype experience homologous net forces. The scaling
Kanatani et al. (1995) shallow foundation, and relationships for the variables contributing to the primary modes of
Meymand (1998) aluminum piles with a SDOF system response are presented in Table 2. By adopting an appropriate
Maugeri et al. (2000) steel structure. geometric scaling factor (l) small-scale models could save cost;
Lu et al. (2004) Adopted soil model: dry however, the precision of the results could be substantially reduced.
Moss et al. (2010) sand, saturated sand, and Therefore, considering the aforementioned specifications of the UTS
reconstituted clayey soil. shaking table, a scaling factor of 1:30 provides the largest achievable
scale model with rational scales, maximum payload, and overturning
Laminar Jafarzadeh and Adopted structural model: moment meeting the facility limitations, which is adopted in the scale
Yanagisawa (1995) steel piles only, concrete model for experimental shaking table tests in this study.
Taylor et al. (1995) piles only, SDOF steel According to Table 2, apart from the geometric scaling that should
Ishimura et al. (1992) structure on shallow be imposed on all components, the required scaled natural frequency
Taylor (1997) foundation, and SDOF steel for the structural models should be 5.83, 2.83, and 2.11 Hz for 5-, 10-,
Tao et al. (1998) structure on concrete piles. and 15-story buildings, respectively. Also, the required scaled shear-
Endo and Komanobe Adopted soil model: dry wave velocity and density of the soil mix should be 36 m=s and
(1995) sand, moist sand, poorly 1,470 kg=m3 , respectively.
Jakrapiyanun (2002) graded river sand, and The input ground motions in this study are represented by a set of
Prasad et al. (2004) saturated sand. real earthquakes defined at the outcropping bedrock. Each test model
Pitilakis et al. (2008) was subjected to two near-field shaking events including Kobe 1995
Chau et al. (2009) and Northridge 1994; two far-field earthquakes including El Centro
Tang et al. (2009) 1940 and Hachinohe 1968; and a sine sweep test. The characteristics of
Turan et al. (2009) the aforementioned earthquakes suggested by the International Asso-
Chen et al. (2010) ciation for Structural Control and Monitoring for benchmark seismic
Lee et al. (2012) studies (Karamodin and Kazemi 2010) are summarized in Table 3. As
Tsukamoto et al. (2012) with the other components of the model, the imposed earthquake
Note: SDOF 5 single degree of freedom. excitations should be scaled as well. Referring to Table 2, although the
model earthquake magnitude remains the same as the prototype, the
time intervals of the original records should be reduced by a factor
this study as shown in Fig. 2. Each frame consisted of three spans of 5.48 (l1=2 , l 5 1=30), which means that the scaled earthquakes
with a total width of 12 m, and the spacing between the frames into contain higher frequencies and shorter durations. The scaled acceler-
the page was 4 m. The natural frequencies of the prototype buildings ation records of the four adopted earthquakes together with the relevant
were 1.05, 0.51, and 0.384 Hz, and the total masses of the prototype frequency content obtained from fast Fourier transform are illustrated in
buildings were 387,000, 671,000, and 953,000 kg for the 5-, 10-, and Figs. 3. In addition, an exponential sine sweep wave with amplitude
15-story buildings, respectively. The characteristics of the prototype of 0:05g, exponential increase rate of 0.5 Hz/min., and frequency
buildings were obtained from the preliminary design of the buildings range of 1–50 Hz was applied to the test models to identify the dy-
following the routine design process according to the Australian namic characteristics of the systems.
Standards AS1170-4 [Standards Association of Australia (SAA)
2007] and AS3600 (SAA 2009a).
The soil medium beneath the structures was a clayey soil with Shaking Table Tests on the Fixed-Base
a shear-wave velocity of 200 m=s and density of 1,470 kg=m3 . The Model Structures
shear-wave velocity of soils in the field can be obtained from a
downhole test, where the soil stiffness properties can be determined by Based on the scaling relationships (Table 2), the model structures were
analyzing direct compressional and shear waves along a borehole down designed to simulate the prototype structures for the shaking table tests.
to about 30 m. Alternatively, the bender element test can be adopted A 3D numerical model was generated, which consisted of horizontal
to measure the shear-wave velocity of the soil samples taken from the steel plates as the floors and four vertical steel plates as the columns.
field. The adopted soil density (1,470 kg=m3 ) can be used as a typical Grade 250 steel plates with a minimum yield stress of 280 MPa and
density for clayey soils as reported by other researchers (e.g., Reza a minimum tensile strength of 410 MPa were adopted in the design,
Tabatabaiefar et al. 2013). The horizontal distances of the soil lateral according to Australian Standard AS/NZS3678 (SAA 2011). The
boundaries and bedrock depth were selected to be 60 and 30 m, re- thickness of the steel plates was determined in the design process
spectively. The buildings rested on a footing that was 1 m thick and after several trial and error cycles to fit the required natural fre-
15 m wide, connected to a 4 3 4 RC pile group with pile diameter and quency and mass of the model structures. The finalized base plate

© ASCE 04014046-3 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. (a) Prototype 15-story building supported by end-bearing pile foundation; (b) prototype 10-story building supported by end-bearing pile
foundation; (c) prototype 5-story building supported by end-bearing pile foundation

Table 2. Scaling Relationships in Terms of the Geometric Scaling in the time domain. In addition, by recording the accelerometers,
Factor (l) which were installed on two edges of the top floor, any possible
Parameter Scaling relationship torsion of the structure during the seismic excitations could be
monitored.
Mass density 1 Initially, a sine sweep test was performed on the structural
Force l3 models to determine the natural frequency of the models. The sine
Stiffness l2 sweep test involves a logarithmic frequency sweep holding
Modulus l a specified acceleration constant at the base of the structure. For
Acceleration 1 the current sine sweep test, by increasing the frequency of the
Shear-wave velocity l1=2 shaking table from 0.1 to 50 Hz, the first resonance between
Time l1=2 the shaking table and the structural model frequencies showed the
Frequency l21=2 fundamental natural frequency of the model. The resulting nat-
Length l ural frequencies of the constructed structural models obtained
Stress l from sine sweep test results were 5.6, 2.85, and 2.19 Hz for the
Strain 1 5-, 10-, and 15-story model structures, respectively, which were
EI l5 in very good agreement with the desired natural frequencies of
the structural models (5.83, 2.83, and 2.11 Hz). In addition, the
estimated value of the structural damping ratio of the constructed
was a 500 3 500 3 10 mm steel plate. The floors consisted of 400 structural model was determined to be approximately 1.1%,
3 400 3 5 mm plates, and four 500 3 40 3 2 mm steel plates were which was obtained from the free vibration lateral displacement
used for the columns. The connections between the columns and floors records of the structural models using Taylor series expansion
were provided using stainless steel metal screws with 2.5 mm diameter (Craig and Kurdila 2006).
and 15 mm length, taking into consideration construction limitations After ensuring the adequacy of the structural model dynamic
and the thickness of the steel plates. After the numerical modeling and characteristics to model the prototype, shaking table tests were
design, the structural models were constructed in-house. The char- performed by applying scaled earthquake acceleration records of
acteristics of the model structures are presented in Table 4, and the Northridge 1994 [Fig. 3(a)], Kobe 1995 [Fig. 3(b)], El Centro
the completed 5-, 10-, and 15-story structural models are illustrated 1940 [Fig. 3(c)], and Hachinohe 1968 [Fig. 3(d)] earthquakes to the
in Fig. 4. fixed-base structural models. The results in terms of the maximum
The first stage of the shaking table tests was carried out under lateral deflections are presented subsequently. To determine the
the fixed-base condition in which the constructed structures were lateral deflections, the movement of the shaking table was subtracted
directly fixed on top of the shaking table (Fig. 4). The arrange- from the story movements. Therefore, all of the records were relative
ment of the installed displacement transducers and accelerome- to the base movements. The results presented here are based on the
ters on the model structures are presented in Table 4, which were lateral deformation of each story when the maximum deflection at the
used to monitor the dynamic response of the structures and to top level occurs. This approach gives a more reasonable pattern of
primarily measure the structural lateral displacements. The the structural deformation in comparison with the approach in which
recorded accelerations can be used to check the consistency and the maximum absolute story deformations are recorded irrespective of
accuracy of the obtained displacements through a double integration occurrence time (Hokmabadi et al. 2012b).

© ASCE 04014046-4 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Table 3. Earthquake Base Motions Used
Earthquake Country Year Peak ground acceleration (g) Magnitude (R) Duration (s) Type Hypocentral distancea (km)
Northridge United States 1994 0.843 6.7 30.0 Near field 9.2
Kobe Japan 1995 0.833 6.8 56.0 Near field 7.4
El Centro United States 1940 0.349 6.9 56.5 Far field 15.69
Hachinohe Japan 1968 0.229 7.5 36.0 Far field 14.1
a
Obtained from Pacific Earthquake Engineering Research Center (PEER 2012).

required bearing capacity underneath the structural model, meeting


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

the scale modeling criterion of the shear-wave velocity. Without


providing enough shear strength for the structural model foundation,
the soil underneath may experience failure or excessive deformation
while the testing process is being undertaken.
To find the most appropriate soil mix for the testing program,
a synthetic clay mixture was designed using Q38 kaolinite clay,
Active Bond 23 bentonite, Class F fly ash, lime, and water as the
components, and a bender element test was performed to measure
the shear-wave velocity of the mix over the cure age (Fatahi et al.
2013). To carry out the bender element tests, the soil specimens were
placed between the bender elements, and the shear-wave velocity of
each soil specimen was obtained by measuring the time required for
the wave to travel between two bender elements using geotechnical
digital systems (GDS) bender element control software running on
a PC. The adopted system has a data acquisition speed of 2,000,000
samples per second, 16 bit as the resolution of data acquisition, and
a connection to the control box through a universal serial bus link. In
this study, the propagated shear-wave type was sine wave with
amplitude of 10 V and a period of 1 s. Several mixtures were ex-
amined, and finally a desired soil mix that consisted of 60% Q38
kaolinite clay, 20% Active Bond 23 bentonite, 20% Class F fly ash
and lime, and 120% water (percentage of the dry mix) produced the
required scaled shear-wave velocity of 36 m=s on the second day of
its cure age. In addition, unconfined compression tests were per-
formed in accordance with Australian Standard AS5101-4 (SAA
2008) to ensure the adequacy of the undrained shear strength of the
proposed soil mix. Table 5 summarizes the soil mix properties on the
second day of its cure age that were adopted in this study. Ac-
cordingly, the soil density on the second day was determined to be
1,450 kg=m3 , which was almost equal to the required prototype soil
density (1,470 kg=m3 ). Therefore, the designed soil mix possessed
the required dynamic similitude characteristics.

Design of the Laminar Soil Container


Fig. 3. Adopted shaking events in this study: (a) scaled Northridge When conducting earthquake model tests, one of the main con-
earthquake; (b) scaled Kobe earthquake; (c) scaled El Centro earth- cerns is the boundary effects created by the artificial boundaries of
quake; (d) scaled Hachinohe earthquake a model container. The premier task of the soil container is to hold
the soil in place during imposed excitation and to provide con-
finement. However, the ideal soil container should simulate free-
field soil behavior as it exists in the prototype by minimizing the
Shaking Table Test on the Structures with Piles boundary effects. The key parameter in the design of the soil
container is to satisfy the same dynamic shear stiffness between the
soil container and the adjacent soil deposit to achieve the strain
Soil Mix
similitude. Moreover, this criterion leads to minimizing the in-
According to the scaling relationships (Table 2), the model soil should teraction between the soil and the container during the shaking
possess a shear-wave velocity of 36 m=s and density of 1,470 kg=m3 table test, and avoiding the generation of primary (compressive)
to provide the desired dynamic similitude between the prototype and waves from the end walls. According to Zeng and Schofield (1996),
the model during the shaking table tests. Previous researchers (e.g., the same dynamic shear stiffness means that the soil container should
Meymand 1998; Turan 2009; Moss et al. 2010) have reported that have the same natural frequency and deflection as the soil deposit in
a reconstituted soil would not be able to satisfy the competing scale the model container.
modeling criterion of shear-wave velocity with enough bearing ca- A 3D explicit finite difference–based program for Fast La-
pacity for the foundation in shaking table tests, while a synthetic clay grangian Analysis of Continua, FLAC3D, version 4.0 (Itasca 2009)
mix can provide adequate undrained shear strength to mobilize the was used to design the laminar soil container for the shaking table

© ASCE 04014046-5 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Table 4. Characteristics of the Model Structures
Location of displacement Location of
Model structure Height (m) Length (m) Width (m) Natural frequency (Hz) Mass (kg) transducers accelerometers
S5 (5-story) 0.5 0.40 0.40 5.6 43 1, 3, 5 1, 2, 3, 4, 5
S10 (10-story) 1.0 0.40 0.40 2.85 74.6 1, 3, 5, 7, 9, 10 1, 3, 5, 7, 8, 9, 10
S15 (15-story) 1.5 0.40 0.40 2.19 106 3, 5, 7, 11, 13, 15 3, 5, 7, 9, 11, 13, 15
Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Model structures for shaking table tests: (a) 15-story model structure; (b) 10-story model structure; (c) 5-story model structure

  
tests. A fully nonlinear time-history dynamic analysis can be per- Gs ¼ Gmax 1 þ g gref (1)
formed using FLAC3D to simulate the realistic dynamic behavior
of the soil and the container under seismic excitations. Initially, the
extensive uniform soil layer was modeled to predict the response of where Gs 5 shear modulus of the soil; Gmax 5 maximum shear
the soil deposit without being interrupted by the artificial boundaries, modulus of the soil in low strains; g 5 cyclic shear strain; and gref
which was used as a benchmark to design the laminar soil container. 5 Hardin-Drnevich constant. A value of gref 5 0:234, giving a co-
Solid elements were used to model the soil deposit, and free-field efficient of determination (R2 ) equal to 0.91, produces the best match
boundary conditions were applied. The nonlinearity of the soil to the backbone curves suggested by Sun et al. (1988) for fine-grained
medium plays a very important role in the seismic behavior of the soils as illustrated in Fig. 5, and thus was adopted in this study. The
soil-pile-structure interaction (Fatahi and Tabatabaiefar 2013; maximum deformations of the soil deposit under the four earthquakes
Hokmabadi et al. 2011; Hokmabadi et al. 2012a, Kim and Roesset are shown in Fig. 6. In addition, applying the sine sweep wave, the
2004; Maheshwari and Sarkar 2011). The built-in tangent modulus natural frequency of the soil deposit was determined to be equal to
function developed by Hardin and Drnevich (1972) was adopted to 10 Hz, which was used in the primary design of the laminar soil
implement hysteretic damping of the soil, which represented the container.
variation of the shear modulus reduction factor and damping ratio Based on the adopted geometric scaling factor (1:30) and
with cyclic shear strain of the soil. This model is defined as follows: allowing a further 10 mm on each side for construction purposes

© ASCE 04014046-6 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


similar to Prasad et al. (2004), the final length, width, and depth of during the test. This method has been also used by other researchers
the laminar soil container were selected to be 2.10, 1.30, and (e.g., Meymand 1998; Taylor 1997) and provides friction between
1.10 m, respectively. The laminar soil container used consisted of a the timber base plate (as bedrock) and the in situ soil to aid in the
rectangular laminar box that was made of aluminum rectangular transmission of shear waves, ensuring negligible relative slip
hollow section frames and separated by rubber layers. The aluminum between the soil and the bottom surface of the container. Moreover,
frames provide lateral confinement of the soil, while the rubber the side walls, which are parallel to the shaking direction, were
layers allow the container to deform in a shear beam manner. The lubricated to be frictionless in order to avoid the generation of shear
numerical model components of the container in FLAC3D are shown stress on the side walls. Fig. 7(b) shows details of the constructed
in Fig. 7(a). The predicted maximum deformations of the empty laminar soil container adopted in this study.
container and the container filled with the desired soil mix were The constructed container was exposed to the same shaking
compared with the free-field deformation of the soil deposit, as events, and the results in terms of the maximum lateral deformation
Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

shown in Fig. 6. The soil container used possessed natural fre- are presented in Fig. 6. The measurements were in good agreement
quencies of 10.8 and 10.4 Hz for the empty and filled conditions, with the predictions, which confirmed the dynamic shear stiffness
respectively, which is in a good agreement with the predicted value similarity between the laminar container and the ideal free-field
(10 Hz). behavior of the soil deposit. As previously mentioned, the soil
The soil container was constructed in-house and the required deposit had a fully nonlinear behavior and its stiffness and strength
rubber sections were manufactured specifically for this project. changed under cyclic loading; however, the stiffness of the con-
The container was secured on the shaking table using eight M38 tainer was constant during the tests. As a result, during excitations
bolts passing through the provided holes. The internal surface of in the lower shear strain range (smaller deformation), the soil
the soil container was covered and sealed with two layers of black dynamic shear stiffness is much closer to the container in com-
plastic sheeting. Following Gohl and Finn (1987) and Valsangkar parison with earthquakes with higher intensity in which more
et al. (1991), 25-mm-thick absorbing layers of polystyrene foam reduction in the initial shear stiffness of the soil is generated
sheets were installed at the end walls of the soil container to (Fig. 5), resulting in more disparities in terms of the maximum
simulate viscous boundaries in the free-field condition and min- lateral deformations.
imize the reflection of the outward propagating waves back into the After ensuring the adequacy of the dynamic characteristics of the
model. In addition, a layer of well-graded gravel was glued (using constructed laminar soil container, 2 m3 of the designed soil mix
strong Megapoxy 69 glue manufactured by Vivacity Engineering [Mix C: 60% kaolinite, 20% bentonite, 20% Class F fly ash and lime,
PTY Ltd., New South Wales, Australia) to the bottom of the soil and water (120% of the dry mix)] were produced and placed into the
container to create a rough interface between the soil and the base laminar soil container. As previously explained, the desired soil mix
acquired the required stiffness, and consequently the shear-wave
velocity, after 2 days of curing. As a result, the entire mixing process
and filling the laminar soil container were completed in 1 day, and
Table 5. Properties of the Soil Mix on the Second Day of Curing Adopted the soil mix inside the container was left to be cured for 2 days while
in the 3D Numerical Model the surface of the soil container was covered and sealed. The soil mix
Soil property Value was poured into the container from a constant height to provide
3 uniformity without imposing any extra compaction or consolidation.
Mass density, r (kg=m ) 1,450
The adopted soil mix provided sufficient workability for this
Shear-wave velocity, Vs (m=s) 36
method. During the soil mixing process, 10 cylindrical soil samples
Maximum shear modulus, Gmax (kPa) 1,776
of D 5 50 mm and h 5 100 mm were taken from the soil mix for
Undrained shear strength, Cu (kPa) 3.1
quality control of the mix. On the second day, the same excitation
Undrained friction angle, wu (degree) 0
events were applied to the filled soil container and the results were
Plasticity index, PI (percentage) 42
compared with the predicted values (Fig. 6).

Fig. 5. Adopted fitting curve for fine-grained soil in this study: (a) relationships between G=Gmax and cyclic shear strain; (b) relationships between the
damping ratio and cyclic shear strain (data from Sun et al. 1988)

© ASCE 04014046-7 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Three-dimensional numerical predictions versus experimental measurements of the maximum lateral deformation of the soil container under
the influence of the following earthquakes: (a) Northridge; (b) Kobe; (c) El Centro; (d) Hachinohe

Fig. 7. (a) Numerical grid and model components of the laminar soil container in FLAC3D; (b) components of the constructed laminar soil container on
the shaking table

© ASCE 04014046-8 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Model Pile Foundation intended to respond in the elastic range (this assumption was con-
firmed numerically), this criterion was achieved by scaling the
Similar to the model structure, the model piles should be subjected to
flexural rigidity (EI) of the piles following previous studies as given
the competing scale model criteria. To achieve a successful model pile
in Table 1 (l5 , l 5 1=30). In addition, this ensured that the yielding
design, the principal governing factors of pile response such as the
point of the model pile was equal to or greater than the scaled pro-
slenderness ratio L=d, moment-curvature relationship, flexural stiff-
totype. Furthermore, by consistently scaling the stiffness of the soil
ness EI, relative soil-pile stiffness, yielding behavior mechanism, and
and pile, the relative soil-pile stiffness parameter will inevitably be
natural frequency of vibration should be addressed (Meymand 1998).
satisfied. Accordingly, the flexural stiffness of the model piles (EI)
By adopting the geometric similitude, the overall pile slenderness and
and the maximum shear modulus (Gmax ) of the soil mix should be
relative contact surface area would be preserved in the model. This
equal to 96:6 N×m2 (Table 6), and 1,776 kPa (Table 5), respectively.
also guarantees that the pile group relative spacing and consequent
Consequently, the relative soil-pile stiffness, as defined by Stewart
Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

group interaction would be replicated at the model scale. Thus, by


et al. (1994), is equal to 2:1 3 1025 for very small shear strains
considering the geometric scaling factor (l) of 1:30 in this study, the
(g , 0:001%). Obviously, the actual relative soil-pile stiffness is
model piles should have a diameter of 40 mm and L=d ratio of 25.
altered during shaking excitations because of the changes in the
The moment-curvature relationship criterion represents the pile
generated shear stiffness in the soil elements according to Fig. 5.
response to lateral loading, which is a function of the flexural rigidity
Previous researchers have used various types of materials such as
and yielding behavior. In the current study, because the piles were
aluminum tubes, steel bars, and RC to build a model pile (e.g., Tao et al.
1998; Chau et al. 2009; Bao et al. 2012). Considering the selected
Table 6. Characteristics of the Model Pile Built from Polyethylene scaling factor in this study (l 5 1=30), and in turn the required stiffness
Pressure Pipe and yielding stress for the model piles, a commercial polyethylene high-
Pile property Value pressure pipe with a standard dimension ratio of 7.4, according to
Australian Standard AS/NZS4130 (SAA 2009b), was the selected
Outer diameter (mm) 40
candidate, which falls in the range of acceptable criteria with 5% de-
Wall thickness (mm) 5.5
viation from the target value for EI. Moreover, polyethylene pipes can
Cross-sectional area (mm2 ) 5:78 3 102
tolerate large deformations prior to the yielding point without any brittle
Moment of inertia (mm4 ) 8:33 3 104
failure. The characteristics of the model pile used in this study are
Young’s modulus (MPa) 1:16 3 103
summarized in Table 6.
Flexural stiffness, EI (N m2 ) 96.6
Relative soil-pile stiffness 2:1 3 1025
Density (kg=m3 ) 955 Testing Procedure
Poisson’s ratio 0.4
The second stage of the shaking table tests was to investigate
Flexural yield stress (MPa) 32
the influence of the soil-pile-structure interaction on the seismic

Fig. 8. Final setup of the shaking table tests for the 10-story model structure with an end-bearing pile foundation

© ASCE 04014046-9 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Fig. 9. (a) Plan of the pile group foundation for the shaking table tests; (b) location of the installed strain gauges on the pile elements

Fig. 10. Recorded maximum lateral deflection of the 5-story model structure from the shaking table tests under the influence of the following
earthquakes: (a) Northridge; (b) Kobe; (c) El Centro; (d) Hachinohe

response of the superstructure in comparison with the conducted possible sliding between the piles and base plate, the pile tips
fixed-base tests. As mentioned previously, commercial poly- were equipped with bolts that were driven into the wooden base
ethylene pressure pipe was used to build the model piles. Wooden plate during the installation. The model piles were driven into the
tips were fitted to the model piles to provide a closed-end con- soil through a 150-mm-high wooden template to ensure location
dition. Also, to compel the end-bearing behavior and prevent any and verticality. Moreover, using a template during the installation

© ASCE 04014046-10 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. Recorded maximum lateral deflection of the 10-story model structure from the shaking table tests under the influence of the following
earthquakes: (a) Northridge; (b) Kobe; (c) El Centro; (d) Hachinohe

process helped to achieve full connection between the piles and levels of the structural model for the 10-story end-bearing pile foun-
the surrounding soil without generating any gaps as a result of the dation system on the shaking table, is presented in Fig. 8.
installation process. The template was constructed with special Strain gauges were installed on the pile elements, as shown in Fig. 9.
cutouts to accommodate a few millimeters of clearance for piles Piles 1, 6, 11, and 16 were instrumented and the strain—and, in turn, the
with external strain gauges to prevent any possible damage to the flexural moments—were measured during the tests at three points along
strain gauges during installation. The size of the cutouts was the piles. Because of the small diameter of the model piles (inner di-
smaller than the diameter of the model piles to avoid lateral ameter 5 29 mm), the closed-end connections at both sides of the mod-
movement of piles during installation. el piles, and the construction limitations, the strain gauges were installed
After installation of the model piles, the template was removed on the outer surface of the model piles. The installed strain gauges were
and a steel plate (simulating the foundation) with prefabricated holes sealed with a waterproof cover.
was fitted over the group. Sixteen M12 bolts were used to provide Similar shaking events, including scaled Kobe 1995, Northridge
a fixed connection between the pile heads and the steel plate. The 1994, El Centro 1940, and Hachinohe 1968 earthquakes, together with
required nuts were fixed to the pile top with strong glue and steel the sine sweep test were applied to the end-bearing pile foundation
rings before the test and the strength and capability of this connec- systems. The natural frequencies of the soil-pile-structure model from
tion technique was examined successfully. Then, the model struc- the performed sine sweep test were measured to be 5.2, 2.6, and 1.93 Hz
tures were suspended from an overhead crane and connected to the for the 5-, 10-, and 15-story model structures, respectively. The results
steel plate at prelocated connections similar to those used in the of the conducted shaking table tests under the influence of the four
fixed-base case. scaled earthquake acceleration records in terms of the maximum lateral
After installing all of the components of the system, including the deflections for the 5-, 10-, and 15-story model structures are presented
container, soil, piles, and superstructure, the same arrangement of dis- in Figs. 10–12. A discussion of the experimental results is presented
placement transducers and accelerometers was used in the model subsequently.
structures to make it comparable with the fixed-base models as pre-
sented in Table 4. In addition, displacement transducers were placed
on the level of the base plate of the structure (simulating the foundation) Results and Discussion
to determine the vertical displacements and rocking angle of the
structure during the testing process. The final setup of the tests, in- According to Fig. 10, the maximum lateral deformation of the 5-
cluding the displacement transducers and accelerometers at various story structure supported by the end-bearing pile foundation

© ASCE 04014046-11 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 12. Recorded maximum lateral deflection of the 15-story model structure from the shaking table tests under the influence of the following
earthquakes: (a) Northridge; (b) Kobe; (c) El Centro; (d) Hachinohe

Table 7. Maximum Rocking Angle of the Base Plate Obtained from Shaking Table Tests
Rocking angle of the foundation
Model structure (5-story) Model structure (10-story) Model structure (15-story)
Scaled earthquake
acceleration record Fixed base Pile foundation (degrees) Fixed base Pile foundation (degrees) Fixed base Pile foundation (degrees)
Northridge 0 0.18 0 0.19 0 0.22
Kobe 0 0.42 0 0.15 0 0.09
El Centro 0 0.10 0 0.13 0 0.14
Hachinohe 0 0.11 0 0.12 0 0.20

increases by 12, 37, 8, and 14% under the influence of the Northridge alter the response of building frames under seismic excitation. This is
1994, Kobe 1995, El Centro 1940, and Hachinohe 1968 earth- because of the fact that the natural period of the system lies in the
quakes, respectively. For the 10-story structure (Fig. 11) and 15- long period region of the response spectrum curve, and the dis-
story structure (Fig. 12), the maximum lateral deformation increases placement response tends to increase while inertia force tends to
by up to 16 and 26%, respectively, in comparison with the fixed-base decrease.
structures. For example, the lateral deformation of the 5-, 10-, and The rocking component plays an important role in the lateral
15-story buildings subjected to the Northridge 1994 earthquake deformation of the superstructure. According to Kramer (1996),
increased by 12, 16, and 18%, respectively, as a result of the SSPSI, relative lateral structural displacements under the influence of the
in comparison with the fixed-base condition. In the meantime, the soil-structure interaction consist of rocking and distortion compo-
natural frequency of the system was reduced because of the seismic nents. According to the recorded maximum rocking angle of the
soil-pile-structure interaction (from 5.6 to 5.2 Hz for the 5-story foundation as given in Table 7, in the 5-story model structure, on
structure; 2.85 to 2.60 Hz for the 10-story structure; and 2.19 to average 16.9% of the maximum lateral deflections were a result of
1.93 Hz for the 15-story structure). Therefore, such decreases in the the rocking component, while 83.1% took place because of the
natural frequency (increases in the natural period) to some extent distortion component. In the 10-story model structure, on average

© ASCE 04014046-12 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 13. Recorded bending moment distribution along Pile 1 supporting 5-, 10-, and 15-story model structures under the influence of the following
earthquakes: (a) Northridge; (b) Kobe; (c) El Centro; (d) Hachinohe

18.8% of the maximum lateral deflections were a result of the the top of the fixed-base 10-story model was measured to be 9.02 mm
rocking component, while 81.2% took place because of the dis- because of the distortion component, while the maximum lateral
tortion component. These values for the 15-story structures were deflection at the top of the same structure supported by the end-
19.7 and 80.3%, respectively. For example, under the influence of bearing pile foundation was 10.22 mm, with 2.35 mm of this value
the 1940 El Centro earthquake, the maximum lateral deflection at being a result of the rocking component and 7.9 mm taking place

© ASCE 04014046-13 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 14. Corresponding interstory drifts of the 5-story model structure from the shaking table tests under the influence of the following earthquakes: (a)
Northridge; (b) Kobe; (c) El Centro; (d) Hachinohe

because of the distortion component. In the pile foundation cases, moment distributions along the piles were obtained from the in-
rocking occurred as a result of the axial deformation of the pile stalled strain gauges during the shaking excitations, and linear in-
elements together with the deformation of the surrounding soil terrelation was adopted between the recorded points to plot the
elements. The area replacement ratio of the pile group was 8% in this bending moment graph versus depth. The distribution of the moment
study, and as a result the piles attract significant axial forces. The area amplitude along the pile showed that the bending moments were
replacement ratio is defined as the cross-sectional area of the pile large at the top of the piles while small bending was generated at the
divided by the tributary area for each pile. The same ratio can also be tips. In addition, by increasing the height of the model structure more
calculated using the cross-sectional area of all piles divided by the bending moments were generated in the pile elements. This is be-
area beneath the pile cap. cause of the fact that the 15-story model structure, as a result of
Comparing the three model structures, the increase in the its larger mass, attracted more inertial force from the same seismic
maximum lateral deflection was more severe for the 15-story excitation in comparison with the 10- or 5-story model structures. As
model structure in comparison with the 10- and 5-story model a result, extra lateral forces and flexural moments were induced in the
structures. This is because of the fact that the rocking component pile foundation system for the 15-story building. As explained
results in more deformation as the height of the structure increases. previously, for the case of the 5-story model structure under the 1995
The influence of SSPSI on the 5-story model structure was con- Kobe earthquake, hitting the structure (with the modified dynamic
siderably higher under the 1995 Kobe earthquake in comparison properties as a result of the SSPSI) close to its natural frequency
with the other earthquakes. This can be a result of the nature of the resulted in attraction of extra energy, which in turn generated extra
imposed seismic motion and its frequency content, resulting in flexural moments in the pile elements during shaking excitations, as
attraction of extra energy by the superstructure and getting close to observed in Fig. 10.
the resonant conditions. This phenomenon is similar to what oc- The corresponding interstory drifts of the model structures were
curred in the lake zone of Mexico City during the 1985 Mexico calculated (Figs. 14–16) using the following equation based on
City earthquake, in which many intermediate height structures Australian standard AS1170-4 (SAA 2007):
with natural periods close to the dominant frequency of the
earthquake (recorded on the ground surface) experienced the most Dði, i þ 1Þ ¼ ðdiþ1 -di Þ=h (2)
damage (Nghiem 2009; Mendoza and Romo 1989).
The measured moment distribution along Pile 1 (Fig. 9), sup- where Dði, i 1 1Þ 5 drift between the (i) and (i 1 1) levels; di11
porting 5-, 10-, and 15-story model structures subjected to the four 5 deflection at the (i 1 1) level; di 5 deflection at the (i) level; and h
previously mentioned earthquake excitations at the instant of peak 5 story height. The plotted drifts (Figs. 14–16) were obtained from
pile head deflection, are presented in Fig. 13. The recorded bending the measured lateral displacements, using the installed displacement

© ASCE 04014046-14 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 15. Corresponding interstory drifts of the 10-story model structure from the shaking table tests under the influence of the following earthquakes:
(a) Northridge; (b) Kobe; (c) El Centro; (d) Hachinohe

transducers at particular stories (as reported in Table 4) during the were conducted on three types of superstructures: 5-, 10-, and 15-
shaking table tests. As observed, considering the effect of the SSPSI story buildings. For each case, two types of foundations (a fixed-
increases the generated interstory drifts in the model structures during base structure and a structure supported by an end-bearing pile
the applied earthquakes. For example, the maximum recorded foundation in soft soil) were experimentally modeled under 1g
interstory drift of the 15-story fixed-base structure subjected to the conditions. An advanced laminar soil container was designed and
1940 El Centro earthquake was measured to be 1.05%, while the tested using 3D numerical modeling that account for the nonlinear
corresponding value for the end-bearing pile foundation case was behavior of the soil.
1.65%. In other words, the SSPSI induces 56% increase in the The details of the improved physical modeling techniques used
recorded maximum interstory drifts in this case. For 10- and 5-story in this study, such as designing the soil mix and laminar soil
model structures subjected to the 1940 El Centro earthquake, con- container, have been explained such that they can be used by other
sidering the effect of the SSPSI induces 14 and 3% increase in the researchers to achieve more accurate simulations of soil-structure
recorded interstory drifts, respectively. This increase in the interstory interactions in 1g shaking table tests. By comparing the experi-
drifts of the superstructures as a result of the SSPSI can shift the mental results, it can be concluded that considering the effects of
performance level of the structures toward near-collapse or col- the SSPSI can alter the dynamic characteristics of the soil-pile-
lapse levels (BSSC 1997), which is of crucial importance in the structure system. In addition, the lateral deflections of structures
performance-based design of the structures, as mentioned in the sitting on end-bearing pile foundations were amplified in com-
Introduction. Therefore, the amplified values of the interstory drifts parison with the fixed-base model. Generally, this amplification,
as a result of the SSPSI should be controlled and minimized for which is mainly a result of the rocking component, is more severe
both structural and nonstructural elements. The extent that the for taller buildings, considering the range of the buildings in-
presence of the pile foundation influences the structural seismic vestigated in this study. Moreover, by increasing the height of the
response depends on the characteristics of the superstructure, such model structure, more bending moments were generated in the pile
as the structural height. elements. This is because of the fact that the 15-story model
structure, as a result of its larger mass, attracts more inertial force
from the same seismic excitation in comparison with the 10- or 5-
Conclusions story model structures.
Consequently, the SSPSI can increase the lateral deflection, and
In this study, in order to investigate the effects of the SSPSI on the in turn the interstory drifts, of structures sitting on soft soil. This
seismic response of superstructures, a series of shaking table tests increase in the interstory drifts of the structure can change the

© ASCE 04014046-15 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 16. Corresponding interstory drifts of the 15-story model structure from the shaking table tests under the influence of the following earthquakes:
(a) Northridge; (b) Kobe; (c) El Centro; (d) Hachinohe

performance level of the structure to near-collapse or even collapse Endo, O., and Komanobe, K. (1995). “Single and multi directional shaking
levels in performance-based design methods. Therefore, ignoring table tests on sand liquefaction.” Proc., 1st Int. Conf. on Earthquake
the real deformability of the soil-pile system may affect the predicted Geotechnical Engineering, Balkema, Rotterdam, Netherlands, 675–680.
damage level of structural and nonstructural elements as well as the Fatahi, B., Fatahi, B., Le, T. M., and Khabbaz, H. (2013). “Small-strain
properties of soft clay treated with fibre and cement.” Geosynth. Int.,
lateral load-carrying mechanism of soil-structure systems during an
20(4), 286–300.
earthquake. Fatahi, B., and Tabatabaiefar, S. (2013). “Fully nonlinear versus equivalent
linear computation method for seismic analysis of midrise buildings on
soft soils.” Int. J. Geomech., 10.1061/(ASCE)GM.1943-5622.0000354,
References 04014016.
Finn, W. D. L. (2005). “A study of piles during earthquakes: Issues of
ASCE. (2010). “Minimum design loads for buildings and other structures.” design and analysis.” Bull. Earthquake Eng., 3(2), 141–234.
ASCE 7-10, Reston, VA. Gohl, W. B., and Finn, W. D. L. (1987). “Seismic response of single piles in
Bao, Y., Ye, G., Ye, B., and Zhang, F. (2012). “Seismic evaluation of soil– shaking table studies.” Proc., 5th Canadian Conf. on Earthquake En-
foundation–superstructure system considering geometry and material gineering, Canadian Association for Earthquake Engineering, Ottawa,
nonlinearities of both soils and structures.” Soils Found., 52(2), 435–444.
257–278. Guin, J., and Banerjee, P. K. (1998). “Coupled soil-pile-structure interaction
Bathurst, R. J., Keshavarz, A., Zarnani, S., and Take, W. A. (2007). “A analysis under seismic excitation.” J. Struct. Eng., 10.1061/(ASCE)
simple displacement model for response analysis of EPS geofoam 0733-9445(1998)124:4(434), 434–444.
seismic buffers.” Soil Dyn. Earthquake Eng., 27(4), 344–353. Ha, I.-S., Olson, S. M., Seo, M.-W., and Kim, M.-M. (2011). “Evaluation
Building Seismic Safety Council (BSSC). (1997). “NEHRP guidelines of reliquefaction resistance using shaking table tests.” Soil Dyn.
for the seismic rehabilitation of buildings, 1997 edition, Part 1: Provi- Earthquake Eng., 31(4), 682–691.
sions and Part 2: Commentary.” FEMA 273/274, FEMA, Washington, Han, Y. (2002). “Seismic response of tall building considering soil-pile-
DC. structure interaction.” Earthquake Eng. Eng. Vib., 1(1), 57–65.
Chau, K. T., Shen, C. Y., and Guo, X. (2009). “Nonlinear seismic soil-pile- Hardin, B. O., and Drnevich, V. P. (1972). “Shear modulus and damping in
structure interactions: Shaking table tests and FEM analyses.” Soil. Dyn. soils: Design equations and curves.” J. Soil Mech. and Found. Div.,
Earthquake Eng., 29(2), 300–310. 98(7), 667–692.
Chen, J., Shi, X., and Li, J. (2010). “Shaking table test of utility tunnel under Harris, H. G., and Sabnis, G. M. (1999). Structural modeling and experi-
non-uniform earthquake wave excitation.” Soil Dyn. Earthquake Eng., mental techniques, CRC, Boca Raton, FL.
30(11), 1400–1416. Hokmabadi, A. S., Fakher, A., and Fatahi, B. (2011). “Seismic strain wedge
Craig, R. R. J., and Kurdila, A. J. (2006). Fundamentals of structural dy- model for analysis of single piles under lateral seismic loading.” Aust.
namics, Wiley, Hoboken, NJ. Geomech., 46(1), 31–41.

© ASCE 04014046-16 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Hokmabadi, A. S., Fakher, A., and Fatahi, B. (2012a). “Full scale lateral Pacific Earthquake Engineering Research Center (PEER). (2012). “PEER
behaviour of monopiles in granular marine soils.” Mar. Struct., 29(1), Ground Motion Database.” Univ. of California, Berkeley, CA.
198–210. Phanikanth, V., Choudhury, D., and Reddy, G. (2013). “Behavior of single
Hokmabadi, A. S., Fatahi, B., and Samali, B. (2012b). “Recording inter- pile in liquefied deposits during earthquakes.” Int. J. Geomech., 10.1061/
story drifts of structures in time-history approach for seismic design of (ASCE)GM.1943-5622.0000224, 454–462.
building frames.” Aust. J. Struct. Eng., 13(2), 175–179. Pitilakis, D., Dietz, M., Wood, D. M., Clouteau, D., and Modaressi, A.
Ishimura, K., Ohtsuki, A., Yokoyama, K., and Koyanagi, Y. (1992). “Sway- (2008). “Numerical simulation of dynamic soil-structure interaction
rocking model for simulating nonlinear response of sandy deposit with in shaking table testing.” Soil Dyn. Earthquake Eng., 28(6), 453–467.
structure.” Proc., 10th World Conf. on Earthquake Engineering, Balkema, Prasad, S., Towhata, I., Chandradhara, G., and Nanjundaswamy, P. (2004).
Rotterdam, Netherlands, 1897–1903. “Shaking table tests in earthquake geotechnical engineering.” Curr. Sci.,
Itasca. (2009). FLAC3D version 4.00 fast Lagrangian analysis of continua 87(10), 1398–1404.
in three dimentions, User’s manual, Minneapolis. Rayhani, M., and El Naggar, M. (2008). “Numerical modeling of seismic
Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

Jafarzadeh, F., and Yanagisawa, E. (1995). “Settlement of sand models response of rigid foundation on soft soil.” Int. J. Geomech., 10.1061/
under unidirectional shaking.” Proc., 1st Int. Conf. on Earthquake (ASCE)1532-3641(2008)8:6(336), 336–346.
Geotechnical Engineering, Balkema, Rotterdam, Netherlands, 693– Reza Tabatabaiefar, S. H., Fatahi, B., and Samali, B. (2013). “Seismic be-
698. havior of building frames considering dynamic soil-structure interaction.”
Jakrapiyanun, W. (2002). “Physical modeling of dynamics soil-foundation- Int. J. Geomech., 10.1061/(ASCE)GM.1943-5622.0000231, 409–420.
structure-interaction using a laminar container.” Ph.D. thesis, Univ. of Richards, R., Jr., Elms, D. G., and Budhu, M. (1990). “Dynamic fluidization
California, San Diego. of soils.” Int. J. Geotech. Eng., 10.1061/(ASCE)0733-9410(1990)116:
Karamodin, A. K., and Kazemi, H. H. (2010). “Semi-active control of 5(740), 740–759.
structures using neuro-predictive algorithm for MR dampers.” Struct. Sato, H., Tanaka, Y., Kanatani, M., Tamari, Y., and Sugisawa, M. (1995).
Contr. Health Monit., 17(3), 237–253. “An experimental and numerical study on the behaviour of improved
Kanatani, M., Nishi, K., and Touma, J. (1995). “Large shake table tests on grounds.” Proc., 1st Int. Conf. on Earthquake Geotechnical Engi-
saturated sand layer and numerical simulation by nonlinear analysis neering, Balkema, Rotterdam, Netherlands, 767–772.
method.” Proc., 1st Int. Conf. on Earthquake Geotechnical Engineering, Shelke, A., and Patra, N. (2008). “Effect of arching on uplift capacity of pile
Balkema, Rotterdam, Netherlands, 705–710. groups in sand.” Int. J. Geomech., 10.1061/(ASCE)1532-3641(2008)8:
Kim, Y., and Roesset, J. (2004). “Effect of nonlinear soil behavior on in- 6(347), 347–354.
elastic seismic response of a structure.” Int. J. Geomech., 10.1061/ Small, J., and Zhang, H. (2002). “Behavior of piled raft foundations under
(ASCE)1532-3641(2004)4:2(104), 104–114. lateral and vertical loading.” Int. J. Geomech., 10.1061/(ASCE)1532-
Kramer, S. L. (1996). Geotechnical earthquake engineering, Prentice Hall,
3641(2002)2:1(29), 29–45.
Upper Saddle River, NJ. Standards Association of Australia (SAA). (2007). “Structural design
Langhaar, H. (1951). Dimensional analysis and theory of models, Wiley,
actions—Earthquake actions in Australia.” AS1170-4, North Sydney,
New York.
New South Wales, Australia.
Lee, C.-J., Wei, Y.-C., and Kuo, Y.-C. (2012). “Boundary effects of
Standards Association of Australia (SAA). (2008). “Methods for preparation
a laminar container in centrifuge shaking table tests.” Soil Dyn. Earth-
and testing of stabilised materials, Method 4: Unconfined compressive
quake Eng., 34(1), 37–51.
strength of compacted materials.” AS5101-4, North Sydney, New South
Lu, X., Li, P., Chen, Y., and Chen, B. (2004). “Shaking table model testing
Wales, Australia.
on dynamic soil-structure interaction system.” Proc., 13th World Conf.
Standards Association of Australia (SAA). (2009a). “Concrete structures.”
on Earthquake Eng., Int. Association for Earthquake Engineering (IAEE),
AS3600, North Sydney, New South Wales, Australia.
Tokyo, Paper No. 3231.
Standards Association of Australia (SAA). (2009b). “Polyethylene (PE)
Ma, X. H., Cheng, Y. M., Au, S. K., Cai, Y. Q., and Xu, C. J. (2009).
pipes for pressure applications.” AS/NZS4130, North Sydney, New
“Rocking vibration of a rigid strip footing on saturated soil.” Comput.
Geotech., 36(6), 928–933. South Wales, Australia.
Maheshwari, B., and Sarkar, R. (2011). “Seismic behavior of soil-pile-structure Standards Association of Australia (SAA). (2011). “Structural steel—
interaction in liquefiable soils: Parametric study.” Int. J. Geomech., Hot-rolled plates, floorplates and slabs.” AS/NZS3678, North Sydney,
10.1061/(ASCE)GM.1943-5622.0000087, 335–347. New South Wales, Australia.
Maugeri, M., Musumeci, G., Novità, D., and Taylor, C. A. (2000). “Shaking Stanton, J. F., Banerjee, S., and Hasayen, I. (1998). “Shaking table tests on
table test of failure of a shallow foundation subjected to an eccentric piles.” Final Rep., Research Project Y-2811, Task 26, Prepared for
load.” Soil Dyn. Earthquake Eng., 20(5–8), 435–444. Washington State Transportation Communication, Univ. of Washington,
Mendoza, M., and Romo, M. (1989). “Behavior of building foundations in Seattle.
Mexico City during the 1985 Earthquake: Second stage.” Proc., Lessons Stewart, D. P., Jewell, R. J., and Randolph, M. F. (1994). “Design of piled
Learned from the 1985 Mexico Earthquake, Earthquake Engineering bridge abutments on soft clay for loading from lateral soil movements.”
Research Institute, El Cerrito, CA, 66–70. Geotechnique, 44(2), 277–296.
Meymand, P. J. (1998). “Shaking table scale model tests of nonlinear soil- Sun, J., Golesorkhi, R., and Seed, H. B. (1988). “Dynamic moduli and
pile-superstructure in soft clay.” Ph.D. thesis, Univ. of California, damping ratios for cohesive soils.” Rep. No. UCB/EERC-88/15, Earth-
Berkeley, CA. quake Engineering Research Center, Univ. of California, Berkeley, CA.
Mizuno, H., Iiba, M., and Hirade, T. (1996). “Pile damage during the 1995 Tang, L., Ling, X., Xu, P., Gao, X., and Wang, D. (2009). “Shake table
Hyogoken-Nanbu earthquake in Japan.” Proc., Proc. 11th World Conf. test of soil-pile groups-bridge structure interaction in liquefiable
on Earthquake Engineering, Int. Association for Earthquake Engineering ground.” Earthquake Engineering and Engineering Vibration, Univ.
(IAEE), Tokyo, Paper No. 977. of California, Los Angeles, 1–12.
Moncarz, P., and Krawinkler, H. (1981). “Theory and application of ex- Tao, X., and Kagawa, T., Minowa, C., and Abe, A. (1998). “Verification
perimental model analysis in earthquake engineering.” Rep. No. 50, John of dynamic soil-pile interaction.” Proc., 3rd Conf. on Geotechni-
A. Blume Earthquake Engineering Center, Dept. of Civil and Envi- cal Earthquake Engineering and Soil Dynamics, ASCE, New York,
ronmental Engineering, Stanford Univ., Stanford, CA. 1199–1210.
Moss, R. E., Crosariol, V., and Kuo, S. (2010). “Shake Table Testing to Taylor, C. A., Dar, A. R., and Crewe, A. J. (1995). “Shaking table modelling
Quantify Seismic Soil Structure Interaction of Underground Structures.” of seismic geotechnical problems.” Proc., 10th European Conf. on
Proc., Int. Conf. on Recent Advances in Geotechnical Earthquake Earthquake Engineers, Balkema, Rotterdam, Netherlands, 441–446.
Engineering and Soil Dynamics, Univ. of Missouri, Rolla, MO. Taylor, C. A. (1997). “Large scale shaking tests of geotechnical struc-
Nghiem, H. (2009). “Soil-pile-structure interaction effects on high rises tures.” Earthquake Engineering Research Centre, Univ. of Bristol,
under seismic shaking.” Ph.D. thesis, Univ. of Colorado, Denver. Bristol, U.K.

© ASCE 04014046-17 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046


Tsukamoto, Y., Ishihara, K., Sawada, S., and Fujiwara, S. (2012). “Set- Canadian Conf. on Earthquake Engineering, Canadian Association for
tlement of rigid circular foundations during seismic shaking in shaking Earthquake Engineering, Ottawa, 327–334.
table tests.” Int. J. Geomech., 10.1061/(ASCE)GM.1943-5622.0000153, Yan, L., and Byrne, P. M. (1989). “Application of hydraulic gradient si-
462–470. militude method to small-scale footing tests on sand.” Can. Geotech. J.,
Turan, A. (2009). “Physical modeling of seismic soil-structure interaction of 26(2), 246–259.
embedded structures.” Ph.D. theisis, Univ. of Western Ontario, London, Zen, K., Yamazaki, H., Toriihara, M., and Mori, T. (1992). “Shaking table
ON, Canada. test on liquefaction of artificially cemented sands.” Proc., 10th World
Turan, A., Hinchberger, S., and El Naggar, H. (2009). “Design and com- Conf. on Earthquake Engineering, Balkema, Rotterdam, Netherlands,
missioning of a laminar soil container for use on small shaking tables.” 1417–1420.
Soil Dyn. Earthquake Eng., 29(2), 404–414. Zeng, X., and Schofield, A. N. (1996). “Design and performance of an
Valsangkar, A. J., Dawe, J. L., and Mita, K. A. (1991). “Shake table studies equivalent-shear-beam container for earthquake centrifuge modelling.”
of sesimic response of single partially suppported piles.” Proc., 6th Geotechnique, 46(1), 83–102.
Downloaded from ascelibrary.org by CBRI - Central Building Research Institute on 04/16/18. Copyright ASCE. For personal use only; all rights reserved.

© ASCE 04014046-18 Int. J. Geomech.

Int. J. Geomech., 2015, 15(2): 04014046

Vous aimerez peut-être aussi